Geometric Complementarity in Assembly and ... - ACS Publications

Sep 22, 2016 - Julian J. Holstein,. †. Claudia M. Wandtke,. §. Birger Dittrich,. ‡ and Guido H. Clever*,†. †. Faculty of Chemistry and Chemic...
2 downloads 0 Views 2MB Size
Subscriber access provided by Northern Illinois University

Article

Geometric Complementarity in Assembly and Guest Recognition of a Bent Heteroleptic cis-[Pd2LA2LB2] Coordination Cage Witold Marek Bloch, Yoko Abe, Julian Jakob Holstein, Claudia M. Wandtke, Birger Dittrich, and Guido H. Clever J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.6b08694 • Publication Date (Web): 22 Sep 2016 Downloaded from http://pubs.acs.org on September 22, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Geometric Complementarity in Assembly and Guest Recognition of a Bent Heteroleptic cis-[Pd2LA2LB2] Coordination Cage Witold M. Bloch,† Yoko Abe,† Julian J. Holstein,† Claudia M. Wandtke,§ Birger Dittrich,‡ Guido H. Clever*,† †

Faculty of Chemistry and Chemical Biology, TU Dortmund University, Otto-Hahn-Straße 6, 44227 Dortmund (Germany).

§

Institute for Inorganic Chemistry, Georg-August University Göttingen, Tammannstraße 4, 37077 Göttingen (Germany).



Institute for Inorganic Chemistry and Structural Chemistry, Heinrich-Heine University Düsseldorf, Universitätsstraße 1, 40225 Düsseldorf (Germany). ABSTRACT: Due to the inherent difficulties in achieving a defined and exclusive formation of multi-component assemblies against entropic predisposition, we present the rational assembly of a heteroleptic [Pd2LA2LB2]4+ coordination cage achieved through the geometric complementarity of two carefully designed ligands, LA and LB. With Pd(II) cations as rigid nodes, the pure distinctly angular components readily form homoleptic cages; a [Pd2LA4]4+ strained helical assembly and a [Pd4LB8]8+ box-like structure, both of which were characterized by X-ray analysis. Combined however, the two ligands could be used to cleanly assemble a cis-[Pd2LA2LB2]4+ cage with a bent architecture. The same self-sorted product was also obtained by a quantitative cage-to-cage transformation upon mixing of the two homoleptic cages revealing the [Pd2LA2LB2]4+ assembly as the thermodynamic minimum. The structure of the heteroleptic cage was examined by ESI-MS, COSY, DOSY and NOESY methods, the latter of which pointed towards a cis-conformation of ligands in the assembly. Indeed, DFT calculations revealed the angular ligands and strict Pd(II) geometry strongly favor the cis-[Pd2LA2LB2]4+ species. The robust nature of the cis-[Pd2LA2LB2]4+ cage allowed us to probe the accessibility of its cavity, which could be utilized for shape recognition towards stereoisomeric guests. The ability to directly combine two different backbones in a controlled manner provides a powerful strategy for increasing complexity in the family of [Pd2L4] cages and opens up possibilities of introducing multiple functionalities into a single self-assembled architecture.

INTRODUCTION Inspired by the function and complexity of biological systems, the design of artificial supramolecular host assemblies has become a burgeoning area of study, particularly utilizing the powerful tool of coordination-driven self-assembly from ligand and metal building blocks.1 The adjustable cavity of resulting assemblies has produced host systems capable of acting as chemical sensors,2 drug transporters,3 components of complex systems,4 stabilization media5 and more.1c,6 Much of this however, has been achieved by relatively simple homoleptic assemblies consisting of one type of ligand. For the purpose of approaching greater complexity and functionality, the derivation of strategies to control the arrangement of different ligand entities in a single assembly is an area which has recently received much attention.7 For example, Zheng and Stang’s charge separation,7n Schmittel’s steric constraints7s and Fujita’s side chain-directed approach7r are among some of the proven methods to access both complex and functional heteroleptic metallosupramolecular architectures.8 [M2L4] coordination cages assembled from squareplanar metal cations such as Pd(II), Pt(II), Cu(II) and

Ni(II) and banana-shaped bis-monodentate ligands are a family of robust and diverse molecular hosts with the ability to provide accessible cavities due to their symmetric and spatial arrangement of ligands.9 The properties of the most intensively studied [Pd2L4] cages are often imparted through functionalization of the ligand components, a strategy which has yielded assemblies with attractive properties such as selective guest binding,3c,10 stimuli responsive structural transitions,11 or rearrangements,12 redox properties,13 and biological activity.14 However, there still remains great scope to extend the complexity and functionality of these architecturally simple assemblies by incorporating more than one type of ligand entity into the structure. Controlled formation of heteroleptic [Pd2L2Lʹ2]4+ assemblies from unprotected Pd(II) cations however has seldom been reported; often the combination of a metal ion with two or more distinct ligands leads to uncontrolled statistical mixtures or narcissistic self sorting,15 especially in the case of N-donor coordinated Pd(II).16 Nevertheless, some examples exist: Hooley and co-workers reported a degree of control over the formation of heteroleptic cages by endohedral functionalization of banana-shaped ligands with differing

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

degrees of steric bulk. A 1:1 mixture of a [Pd2L4]4+ and [Pd2L3Lʹ1]4+ species could be observed with a combination of bulky and non-bulky ligands.17 On the other hand, Yoshizawa recently reported a [Pd2L2Lʹ2]4+ cage achieved through a template effect with a C60 guest.18 While these systems nicely showcase different strategies to achieve heteroleptic assemblies, they suffer from the problem that their formation is inevitably linked to an already occupied cavity. Most recently, Crowley and co-workers reported a strategy to achieve controlled formation of [Pd2L2Lʹ2]4+ coordination cages through exploitation of H-bonding between electron-rich 2-amino substituted pyridyl ligand components.19 Intriguingly, the kinetically driven heteroleptic assembly could only be accessed through ligand displacement reactions, rather than from direct ligand assembly or cage-to-cage conversion of their homoleptic counterparts. Apart from these examples, strategies to access coordinatively saturated Pd(II) cage assemblies from at least two distinct ligand components remain scarce. Herein we report the pre-programming of ligand components (LA and LB) with shape complementarity based on a directional bonding approach20 as a robust route to the controlled and template-free formation of a 3component [Pd2LA2LB2]4+ assembly (C3). In contrast with most previously reported multi-component assembly strategies,7 the formation here does not rely upon the combination of different donor sets or implementation of steric bulk, but rather the energy benefit associated with utilizing ligands of a complementary shape, which results in the heteroleptic assembly as the stable, thermodynamic product.

Figure 1. Self-assembly scheme showing the ligands and coordination cages presented in this work; a) self-assembly of a

A

Page 2 of 15

4+

homoleptic [Pd2L 4] cage (C1); b) self-assembly of a homoB 8+ leptic [Pd4L 8] box (C2); c) three-component self-assembly A B 4+ of a heteroleptic [Pd2L 2L 2] cage (C3), achieved either from the individual ligand components and Pd(II) or by (i) mixing C1 and C2 in a 2:1 ratio.

RESULTS AND DISCUSSION Synthesis of ligands LA and LB. Initially, we compared the formation of homo- and heteroleptic self-assemblies based on the two shape complementary ligands LA and LB (Figure 1). Therefore, ligand LA was synthesized by Sonogashira cross-coupling of 2,7-dibromo-10-hexylacridin9(10H)-one with 8-ethynyl-isoquinoline, whilst ligand LB was prepared by a Suzuki coupling of 3,6-dibromo-9,10dimethoxyphenanthrene and 4-pyridineboronic acid pinacol ester. The structures of LA and LB were confirmed using NMR spectroscopy, mass spectrometry, and X-ray crystallography (Supporting Information). We then turned our attention to investigating the respective Pd– mediated homoleptic assemblies of LA and LB. Homoleptic assembly of cages C1 and C2. With regards to LA, we anticipated that the eight inward-pointing isoquinoline donors may have to undergo severe twisting in a [Pd2LA4]4+ assembly. Heating a 2:1 mixture of LA and [Pd(CH3CN)4](BF4)2 in DMSO at 70 °C for 2 h resulted in the quantitative formation of a single product, identified as a [Pd2LA4]4+ cage (C1) by 1H NMR spectroscopy (Figure 2b) and mass spectrometry (Figure S27). The 1H NMR spectrum of C1 revealed a significant upfield shift of several isoquinoline and acridone proton signals of LA (Figure 2a and b), suggesting shielding by a neighboring πsystem due to twisting and dense association of the ligand backbones. Indeed the X-ray structure of C1 revealed that the four ligands (two of which are crystallographically unique) are in a highly twisted conformation, participating in either π – stacking or hydrogen bonding with neighboring backbones, resulting in a helical assembly (Figure 5a). Due to packing effects, C1 contains C2 symmetry in the solid-state, however the number of observed NMR signals indicates that the overall flexibility of the assembly allows for a higher, fourfold symmetry in solution. Noteworthy is the compressed nature of the structure, with a Pd···Pd distance of 15.05 Å. For the Pd–mediated assembly of LB, we expected a different structure due to the para-substituted pyridine donors which create an approximate 60° vector angle with respect to the phenanthrene backbone. Indeed, heating a 2:1 mixture of LB and [Pd(CH3CN)4](BF4)2 in DMSO at 70 °C for 2 h resulted in the quantitative formation of a [Pd4LB8]8+ (C2) cage species (Figure 1b). The ESI HR-MS spectrum was consistent with only one major species, yielding signals for ions of [Pd4LB8+nBF4]8−n+ (n=1−4) (Figure S28). In the 1H NMR spectrum, downfield shifts of pyridyl protons Heʹ and Hdʹ were observed (Figure 2e,d), consistent with Pd(II) complexation. Along with 2D NMR analysis, these observations pointed towards a D4h symmetric M4L8 ‘box’ structure for C2, a topology that has previously been encountered by Fujita and co-workers

ACS Paragon Plus Environment

Page 3 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

using a similar ligand.21 The structure of C2 was confirmed by X-ray crystallography which revealed the expected connectivity (Figure 5b). Due to packing effects the assembly deviates slightly from ideal D4h symmetry, with the opposing Pd···Pd distances measuring 17.44 and 16.73 Å respectively.

1

Figure 2. Partial H NMR (400 MHz/DMSO-d6, 25°C) spectra A 4+ B 8+ showing the self-assembly of [Pd2L 4] (C1), [Pd4L 8] (C2), A B 4+ A A 4+ and [Pd2L 2L 2] (C3); a) L ; b) [Pd2L 4] obtained by heatA A B 4+ ing L with 0.5 equiv of Pd(II); c) [Pd2L 2L 2] obtained by A B heating a 1:1:1 mixture of L , L and Pd(II), or a 2:1 mixture of A 4+ B 8+ B 8+ [Pd2L 4] and [Pd4L 8] , respectively; d) [Pd4L 8] obtained B B from heating L and 0.5 equiv of Pd(II); e) L .

Heteroleptic assembly of cage C3. For the target [Pd2LA2LB2]4+ heteroleptic assembly, molecular modeling suggested that the geometric constraints imposed by the metal and ligand components should behave synergistically to yield only one cage isomer. DFT calculations indicated that compared to the trans-[Pd2LA2LB2]4+ cage (+131.8 kJ/mol) the formation of the cis-[Pd2LA2LB2]4+ cage from its components is significantly more energetically favorable (−65.6 kJ/mol) due to the complementary arrangement of the ligands with respect to the Pd(II) coordination sphere. In addition, cages obeying the stoichiometries [Pd2LA3LB1]4+ and [Pd2LA1LB3]4+ were found to be higher in energy than the cis-isomer (Figure 3). Therefore we expected that the square planar geometry of Pd(II) and the respective backbone angles of LA and LB should strongly favor a cis arrangement of the ligands.

Figure 3. Energy diagram with DFT calculated structures of A B 4+ A B 4+ A B 4+ trans–[Pd2L 2L 2] , cis–[Pd2L 2L 2] , [Pd2L 3L 1] and A B 4+ [Pd2L 1L 3] . The energies of the respective cages were calculated according to equations described in the supporting information. To simplify the calculations, the hexyl chain of A the acridone moiety of L was replaced with a methyl substituent.

Indeed, heating a mixture of LA, LB and [Pd(CH3CN)4](BF4)2 in a 1:1:1 ratio for 2 h at 70 °C gave rise to a single species with a distinct 1H NMR spectrum (Figure 2c). Interestingly, the isoquinoline protons of LA, Hi and Hh, were significantly broadened in the room temperature NMR spectra. Therefore, we performed variable temperature 1H NMR experiments (Figure S16) which revealed a gradual sharpening of all signals including the isoquinoline protons. We then performed a 1H – 1H COSY experiment at 70 °C which allowed complete assignment of the expected set of 14 aromatic proton signals for the heteroleptic [Pd2LA2LB2]4+ (C3) species (Figure S18). In addition, we performed a NOESY experiment at 70 °C in order to assign the important inter-ligand contacts in C3 (Figure 4b, Figure S19). Analysis revealed several evident cross-peaks, particularly between the isoquinoline and acridone protons of LA and the pyridyl protons of LB. Importantly, the observed contacts were in full agreement with the calculated model. DOSY analysis confirmed that all of the proton signals assigned to C3 correspond to the same diffusion coefficient (Figure 4c). Further characterization of the sample by ESI HR-MS yielded a relatively simple spectrum, with the prominent signals assigned to the [Pd2LA2LB2+nBF4]4−n+ species (n = 0, 1) (Figure 4a). Cage-to-cage transformation of C1 and C2 to give C3. Given the rapid and facile assembly of C3 from LA, LB and Pd(II), we next performed experiments to investigate whether C3 is the thermodynamic minimum of a mixture of C1 and C2. In contrast to the assembly from the individual ligands, mixing C1 and C2 in a 2:1 ratio resulted in a rather slow conversion to the heteroleptic species C3, complete after 12 days of heating at 70 °C (Figure S20), presumably due to the requirement of disassembling

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 15

multiple coordination bonds.22 We also investigated the conversion to C3 by addition of the either LA or LB to C2 or C1, respectively. We observed that upon addition of LB to C1, C3 was formed immediately at room temperature (Figure S21). Conversely, upon addition of LA to C2, the system reached equilibrium after 2 days of heating at 70 °C with only 10% of C3 formed (Figure S22). We assume differing strain within these species to be responsible for this effect. For the helical structure of C1, addition of LB provides an opportunity to release strain via disassembly and reassembly to C3, whilst the same energy benefit is presumably not provided for the less strained structure of C2.

A

B

4-n+

Figure 4. a) ESI mass spectrum of [Pd2L 2L 2+nBF4] with n A B 4−n+ = 0, 1 (* = [Pd2L 3L 2+nBF4] , with n = 0−2). b) An expan1 1 sion of the H – H NOESY spectrum of C3 measured at 70 °C. The inter-ligand cross peaks are highlighted in red and indicated on the DFT model of C3 in the inset. c) DOSY spectrum (500 MHz/DMSO, 70°C) of C3. All of the signals assigned to C3 (marked with a circle) correspond to the same -10 2 -1 diffusion coefficient (2.14 × 10 m s , log D = −9.67).

Figure 5. A perspective view of the X-ray crystal structures of a) C1 and b) C2 with counter-ions removed for clarity. c) DFT calculated model of C3. The hexyl chain of C3 was omitted to simplify calculations. The space filling representation is overlaid for each structure (supplementary crystallographic data is found in the Supporting Information and CCDC data sets 1472456, 1489224-1489226). The molar ratios of C1 – C3 associated with the cage-to-cage transformation are included below each struc-

ACS Paragon Plus Environment

Page 5 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

ture. The associated energy benefit as roughly estimated by DFT calculations is discussed in the Supporting Information (Figure SI 31).

Despite numerous attempts, we were not able to obtain single crystals of C3 suitable for X-ray analysis. To obtain further insight into the assembly of C3 and investigate the observed intramolecular self-sorting, we further compared the DFT structures of C1, C2, and C3 (Figures S31 and 5c) revealing that the cage-to-cage transformation of C1 and C2 to C3 should be highly energetically favored (Figure 5 and the SI) This result supports C3 as the thermodynamic minimum of the system. Shape complementary guest binding. It is interesting to note that C3 is the first example of a [Pd2L4] coordination cage with a bent architecture. As the Pd(II) metal centers can serve as anchors for charged molecules,21 we identified that such host architecture may possess a shape specific cavity for guest binding. To test this hypothesis, we performed 1H NMR titrations with a straight and bent-shaped guest; 2,7-napthalene disulfonate (G1) and the 2,6 analogue (G2). In both cases, fast exchange was observed relative to the time scale of the experiment (Figure S23 and S24). We determined the host to guest stoichiometry to be 1:1 by the Job plot method (Figure S26), and further verified this by mass spectrometry (Figure S29 and S30). Furthermore, contacts observed in the NOESY analysis of G1@C3 (Figure S25) revealed that the disulfonate guest is situated between the acridone backbones of the two adjacent LA ligands in C3, most likely stabilized by π – stacking. Therefore, from the 1H NMR titrations (Figure 6a) we calculated the association constant between G1 and C3 and G2 and C3 to be approximately 5200 and 2300 M−1 respectively. This difference can be explained by the shape-complementary fit of G1 relative to the cavity and angular Pd(II) anchors of C3. To support these observations, we calculated the structures of G1@C3 and G2@C3 (Figure 6b and c). A comparison of the minimized energies revealed that G1@C3 is stabilized by 40.7 kJ/mol as compared to isomeric complex G2@C3, which is in accordance with the optimal fit of G1 inside the shape-specific cavity of C3. Interestingly, previous binding studies of G1 and G2 in a [Pd2L4]4+ cage23 with relationally parallel Pd(II) planes showed a stronger binding for G2. Thus, the unusual cavity and angular Pd(II) anchors of C3 create a shape-specific environment with opposite guest binding preference than the previously studied example.

1

1

2

Figure 6. H NMR titrations of C3 with G and G . Circles, diamonds, and triangles represent the shift of protons He’, Hc’ B A (L ) and Hc (L ) respectively. c) and d) show the energy 1 2 minimized structure of G @C3 and G @C3 respectively.

CONCLUSIONS In conclusion, we have presented the self-assembly of two complementary ligands LA and LB in homoleptic and heteroleptic Pd-mediated coordination cages. We have shown that geometric complementarily pre-programmed into ligand components is a robust strategy to achieve a stable heteroleptic cis-[Pd2LA2LB2]4+ cage, thus surmounting the entropic tendency to form a mixture of products. Furthermore, we demonstrated that the heteroleptic architecture can be accessed through multiple self-assembly pathways: direct combination of the ligands with Pd(II), cage-to-cage transformations, and ligand induced cage rearrangements. The latter was found to proceed smoothly only in the case of addition of LB to C1, revealing the possible strain in the C1 helical species as the driving force for this reaction. The cage-to-cage transformations also highlighted an important feature of our system; the thermodynamic stability of the heteroleptic product, allowing us to probe the cavity of C3. The unique shape of C3 and angular Pd(II) anchors indeed provided an accessible cavity which we exploited in the shape recognition on the level of host-guest binding. We think that the implementation of this strategy into the area of [Pd2L4] cages may yield new and unique host systems with a greater control over the incorporation of multiple functionalities and hence fine tuning of the chemistry of the cavity. Current investigations into functionalizing the

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

individual backbones with complementary entities (e.g. electron-donor acceptor system) are underway in our laboratory.

ASSOCIATED CONTENT Supporting Information. Experimental details and further X-ray, NMR, MS and computational data are presented. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author * E-mail: [email protected]

Notes The authors declare no competing financial interest

ACKNOWLEDGMENT We thank Dr. H. Frauendorf (Göttingen) for measuring the ESI mass spectra and Dr. M. John (Gö.) and Prof. Dr. W. Hiller (TU Dortmund) for assistance with NMR measurements and helpful discussions. W.M.B thanks the Alexander von Humboldt Foundation for a postdoctoral fellowship. Y.A. thanks the Japan Society for the Promotion of Science for a fellowship. The Fonds der Chemischen Industrie and the Deutsche Forschungsgemeinschaft (CL 489/2-1) kindly provided financial support. We thank the European Research Council for supporting our further studies in this direction by ERC Consolidator grant 683083 (RAMSES).

REFERENCES 1. a) Fujita, M.; Tominaga, M.; Hori, A.; Therrien, B. Acc. Chem. Res. 2005, 38, 369; b) Northrop, B. H.; Zheng, Y.-R.; Chi, K.-W.; Stang, P. J. Acc. Chem. Res. 2009, 42, 1554; c) Chakrabarty, R.; Mukherjee, P. S.; Stang, P. J. Chem. Rev. 2011, 111, 6810; d) Smulders, M. M. J.; Riddell, I. A.; Browne, C.; Nitschke, J. R. Chem. Soc. Rev. 2013, 42, 1728; e) Hof, F.; Craig, S. L.; Nuckolls, C.; Rebek, J. J. Angew. Chem. Int. Ed. 2002, 41, 1488. 2. a) Ahmad, H.; Hazel, B. W.; Meijer, A. J. H. M.; Thomas, J. A.; Wilkinson, K. A. Chem. Eur. J. 2013, 19, 5081; b) Neelakandan, P. P.; Jimenez, A.; Nitschke, J. R. Chem. Sci. 2014, 5, 908. 3. a) Therrien, B.; Süss-Fink, G.; Govindaswamy, P.; Renfrew, A. K.; Dyson, P. J. Angew. Chem. Int. Ed. 2008, 47, 3773; b) Lewis, J. E. M.; Gavey, E. L.; Cameron, S. A.; Crowley, J. D. Chem. Sci. 2012, 3, 778. 4. a) Inokuma, Y.; Kawano, M.; Fujita, M. Nat Chem 2011, 3, 349; b) Zarra, S.; Wood, D. M.; Roberts, D. A.; Nitschke, J. R. Chem. Soc. Rev. 2015, 44, 419. 5. a) Kawano, M.; Kobayashi, Y.; Ozeki, T.; Fujita, M. J. Am. Chem. Soc. 2006, 128, 6558; b) Fiedler, D.; Bergman, R. G.; Raymond, K. N. Angew. Chem. Int. Ed. 2006, 45, 745; c) Mal, P.; Breiner, B.; Rissanen, K.; Nitschke, J. R. Science 2009, 324, 1697. 6. a) Ward, M. D.; Raithby, P. R. Chem. Soc. Rev. 2013, 42, 1619; b) Cook, T. R.; Stang, P. J. Chem. Rev. 2015, 115, 7001; c) McConnell, A. J.; Wood, C. S.; Neelakandan, P. P.; Nitschke, J. R. Chem. Rev. 2015, 115, 7729; d) Brown, C. J.; Toste, F. D.; Bergman, R. G.; Raymond, K. N. Chem. Rev. 2015, 115, 3012. 7. Reviews: a) De, S.; Mahata, K.; Schmittel, M. Chem. Soc. Rev. 2010, 39, 1555; b) Saha, M. L.; De, S.; Pramanik, S.; Schmittel, M. Chem. Soc. Rev. 2013, 42, 6860; c) Mukherjee, S.; Mukherjee, P. S. Chem. Commun. 2014, 50, 2239; d) Li, H.; Yao, Z.-J.; Liu, D.; Jin, G.-X. Coord. Chem. Rev. 2015, 293–294, 139; e) He, Z.; Jiang, W.; Schalley, C. A. Chem. Soc. Rev. 2015, 44, 779; f) Newkome, G.

Page 6 of 15

R.; Moorefield, C. N. Chem. Soc. Rev. 2015, 44, 3954. Examples: g) Howlader, P.; Das, P.; Zangrando, E.; Mukherjee, P. S. J. Am. Chem. Soc. 2016, 138, 1668; h) Metherell, A. J.; Ward, M. D. Chem. Sci. 2016, 7, 910; i) Ronson, T. K.; Roberts, D. A.; Black, S. P.; Nitschke, J. R. J. Am. Chem. Soc. 2015, 137, 14502; j) Zhang, L.; Lin, Y.-J.; Li, Z.-H.; Jin, G.-X. J. Am. Chem. Soc. 2015, 137, 13670; k) Lee, H.; Noh, T. H.; Jung, O.-S. Angew. Chem. Int. Ed. 2016, 55, 1005; l) Chepelin, O.; Ujma, J.; Barran, P. E.; Lusby, P. J. Angew. Chem. Int. Ed. 2012, 51, 4194; m) Yamanaka, M.; Yamada, Y.; Sei, Y.; Yamaguchi, K.; Kobayashi, K. J. Am. Chem. Soc. 2006, 128, 1531; n) Zheng, Y.-R.; Zhao, Z.; Wang, M.; Ghosh, K.; Pollock, J. B.; Cook, T. R.; Stang, P. J. J. Am. Chem. Soc. 2010, 132, 16873; o) Schmidtendorf, M.; Pape, T.; Hahn, F. E. Angew. Chem. Int. Ed. 2012, 51, 2195; p) Prusty, S.; Krishnaswamy, S.; Bandi, S.; Chandrika, B.; Luo, J.; McIndoe, J. S.; Hanan, G. S.; Chand, D. K. Chem. Eur. J. 2015, 21, 15174; q) García-Simón, C.; Gramage-Doria, R.; Raoufmoghaddam, S.; Parella, T.; Costas, M.; Ribas, X.; Reek, J. N. H. J. Am. Chem. Soc. 2015, 137, 2680; r) Yoshizawa, M.; Nagao, M.; Kumazawa, K.; Fujita, M. J. Organomet. Chem. 2005, 690, 5383; s) Schmittel, M.; Ganz, A. Chem. Commun. 1997, 999; t) Sun, Q.-F.; Sato, S.; Fujita, M. Angew. Chem. Int. Ed. 2014, 53, 13510. 8. a) Kishore, R. S. K.; Paululat, T.; Schmittel, M. Chem. Eur. J. 2006, 12, 8136; b) Schmittel, M.; Kalsani, V.; Michel, C.; Mal, P.; Ammon, H.; Jäckel, F.; Rabe, J. P. Chem. Eur. J. 2007, 13, 6223; c) Osuga, T.; Murase, T.; Fujita, M. Angew. Chem. Int. Ed. 2012, 51, 12199; d) Yan, X.; Cook, T. R.; Wang, P.; Huang, F.; Stang, P. J. Nat Chem 2015, 7, 342; e) Samanta, D.; Shanmugaraju, S.; Joshi, S. A.; Patil, Y. P.; Nethaji, M.; Mukherjee, P. S. Chem. Commun. 2012, 48, 2298; f) Samanta, D.; Mukherjee, P. S. Chem. Commun. 2014, 50, 1595; g) Cook, T. R.; Vajpayee, V.; Lee, M. H.; Stang, P. J.; Chi, K.-W. Acc. Chem. Res. 2013, 46, 2464. 9. a) Han, M.; Engelhard, D. M.; Clever, G. H. Chem. Soc. Rev. 2014, 43, 1848; b) Custelcean, R. Chem. Soc. Rev. 2014, 43, 1813; c) Frank, M.; Johnstone, M. D.; Clever, G. H. Chem. Eur. J. 2016, 22, 14104. 10. a) Johnstone, M. D.; Schwarze, E. K.; Ahrens, J.; Schwarzer, D.; Holstein, J. J.; Dittrich, B.; Pfeffer, F. M.; Clever, G. H. Chem. Eur. J. 2016, 22, 10791; b) Löffler, S.; Lübben, J.; Wuttke, A.; Mata, R. A.; John, M.; Dittrich, B.; Clever, G. H. Chem. Sci. 2016, 7, 4676; c) Zhou, L.-P.; Sun, Q.-F. Chem. Commun. 2015, 51, 16767; d) Kishi, N.; Li, Z.; Yoza, K.; Akita, M.; Yoshizawa, M. J. Am. Chem. Soc. 2011, 133, 11438; e) Liao, P.; Langloss, B. W.; Johnson, A. M.; Knudsen, E. R.; Tham, F. S.; Julian, R. R.; Hooley, R. J. Chem. Commun. 2010, 46, 4932. 11. a) Han, M.; Michel, R.; He, B.; Chen, Y.-S.; Stalke, D.; John, M.; Clever, G. H. Angew. Chem. Int. Ed. 2013, 52, 1319; b) Löffler, S.; Lübben, J.; Krause, L.; Stalke, D.; Dittrich, B.; Clever, G. H. J. Am. Chem. Soc. 2015, 137, 1060. 12. a) Zhu, R.; Lübben, J.; Dittrich, B.; Clever, G. H. Angew. Chem. Int. Ed. 2015, 54, 2796; b) Sekiya, R.; Fukuda, M.; Kuroda, R. J. Am. Chem. Soc. 2012, 134, 10987; c) Bandi, S.; Pal, A. K.; Hanan, G. S.; Chand, D. K. Chem. Eur. J. 2014, 20, 13122. 13. Frank, M.; Hey, J.; Balcioglu, I.; Chen, Y.-S.; Stalke, D.; Suenobu, T.; Fukuzumi, S.; Frauendorf, H.; Clever, G. H. Angew. Chem. Int. Ed. 2013, 52, 10102. 14. a) McNeill, S. M.; Preston, D.; Lewis, J. E. M.; Robert, A.; Knerr-Rupp, K.; Graham, D. O.; Wright, J. R.; Giles, G. I.; Crowley, J. D. Dalton Trans. 2015, 44, 11129; b) Ahmedova, A.; Momekova, D.; Yamashina, M.; Shestakova, P.; Momekov, G.; Akita, M.; Yoshizawa, M. Chem. Asian J. 2016, 11, 474. 15. a) Safont-Sempere, M. M.; Fernández, G.; Würthner, F. Chem. Rev. 2011, 111, 5784; b) Lal Saha, M.; Schmittel, M. Org. Biomol. Chem. 2012, 10, 4651; c) Jiménez, A.; Bilbeisi, R. A.; Ronson, T. K.; Zarra, S.; Woodhead, C.; Nitschke, J. R. Angew. Chem. Int. Ed. 2014, 53, 4556; d) Yan, L.-L.; Tan, C.-H.; Zhang, G.L.; Zhou, L.-P.; Bünzli, J.-C.; Sun, Q.-F. J. Am. Chem. Soc. 2015,

ACS Paragon Plus Environment

Page 7 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

137, 8550; e) Johnson, A. M.; Wiley, C. A.; Young, M. C.; Zhang, X.; Lyon, Y.; Julian, R. R.; Hooley, R. J. Angew. Chem. 2015, 127, 5733. 16. a) Frank, M.; Ahrens, J.; Bejenke, I.; Krick, M.; Schwarzer, D.; Clever, G. H. J. Am. Chem. Soc. 2016, 138, 8279; b) Frank, M.; Krause, L.; Herbst-Irmer, R.; Stalke, D.; Clever, G. H. Dalton Trans. 2014, 43, 4587. 17. Johnson, A. M.; Hooley, R. J. Inorg. Chem. 2011, 50, 4671. 18. Yamashina, M.; Yuki, T.; Sei, Y.; Akita, M.; Yoshizawa, M. Chem. Eur. J. 2015, 21, 4200. 19. Preston, D.; Barnsley, J. E.; Gordon, K. C.; Crowley, J. D. J. Am. Chem. Soc. 2016, 138, 10578. 20. a) Stang, P. J.; Olenyuk, B. Acc. Chem. Res. 1997, 30, 502; b) Fujita, M.; Fujita, N.; Ogura, K.; Yamaguchi, K. Nature 1999, 400,

52; c) Klotzbach, S.; Beuerle, F. Angew. Chem. Int. Ed. 2015, 54, 10356. 21. a) Suzuki, K.; Kawano, M.; Fujita, M. Angew. Chem. 2007, 119, 2877; b) Kumar Chand, D.; Fujita, M.; Biradha, K.; Sakamoto, S.; Yamaguchi, K. Dalton Trans. 2003, 2750. 22. Wang, W.; Wang, Y.-X.; Yang, H.-B. Chem. Soc. Rev. 2016, 45, 2656. 23. Clever, G. H.; Tashiro, S.; Shionoya, M. Angew. Chem. Int. Ed. 2009, 48, 7010; b) Clever, G. H.; Kawamura, W.; Shionoya, M. Inorg. Chem. 2011, 50, 4689.

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 15

Table of Contents artwork

ACS Paragon Plus Environment

8

Page 9 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

87x92mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

64x49mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 10 of 15

Page 11 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

68x55mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

124x183mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 12 of 15

Page 13 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

59x21mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

20x5mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 14 of 15

Page 15 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

93x104mm (300 x 300 DPI)

ACS Paragon Plus Environment